You are on page 1of 21

Bot. Rev.

(2009) 75:271–291
DOI 10.1007/s12229-009-9032-1

Plant Communities of Western Amazonia

Randall W. Myster1,2
1
Biology Department, Oklahoma State University, 900 N. Portland Ave, Oklahoma City,
OK 73107, USA
2
Author for Correspondence; e-mail: rmyster@lternet.edu

Published online: 23 May 2009


# The New York Botanical Garden 2009

Abstract In this review I focus on the socially and ecologically important western
Amazon basin and its associated plant communities. I delineate this vast area as between
the Andes to the west and the confluence of the Amazon and its first major black-water
river, the Rio Negro, to the east. Although scientists have explored here, and local people
have lived here for years, we still have—unfortunately—only a most basic under-
standing of its plant communities. This review is motivated by that lack of knowledge,
and attempts both to add a level of organization to what we do know and to suggest
future avenues of research. I do this by first realizing that plant communities here must
be fundamentally differentiated by the degree of flooding they experience, the kind of
water involved, and how regularly that flooding occurs. Within that context, plant
communities can then be further defined by such characteristics as soil type, micro-
topological relief, and human-induced disturbance regime. After completing the review,
I suggest that the diversity of plant communities, not just the plants themselves, is large
and likely to grow substantially as we sample more and more, that is to say the peak of
the plant community-sampling area curve has not yet been reached. I close with basic
questions to help guide future efforts, ideas on how plant communities could be defined
quantitatively, and a call for more research funding of the Western Amazon.

Resumen En esta revisión que enfoco en la cuenca del Amazonas socialmente y


ecológicamente importante occidental y se ha asociado las comunidades de planta. Delineo
esta área vasta como entre los Andes y la confluencia del Amazonas y su primer río mayor
de negro-agua, el Negro de Rio. Aunque científicos hayan explorado aquí durante años,
nosotros tenemos todavía—desgraciadamente—sólo un comprensión muy básico de su
estructura. Esta revisión es motivada por esa falta del conocimiento, y las tentativas para
agregar un nivel de la organización a lo que sabemos. Hago que por darse cuenta de
primero que esas comunidades de planta aquí deben ser diferenciadas fundamentalmente
por el grado de la inundación ellos experimentan, la clase de agua implicada, y si esa
inundación ocurre con regularidad o no. Dentro de ese contexto, aún más comunidades de
planta entonces fueron definidas. Encontré que la diversidad de comunidades, no apenas
las plantas sí mismos, son grande y probable de crecer substancial como probamos cada
vez mas, eso es el pico de la curva de comunidad-área tiene todavía no fue alcanzado.
Cierro con algunas preguntas básicas y también algunas ideas de mi propio análisis de
Neotropic campos viejos en cómo nosotros quizás definamos las comunidades
cuantitativamente, así como una llamada para más financiación de investigación
272 R.W. Myster

Introduction

The Amazon is the mightiest river in the world, having a discharge of fresh water 4–
5 times greater than that of the next highest river, the Congo. It is also the second
longest river in the world, originating in the foothills of the Andean Mountains of
South America and running east into the Atlantic Ocean. It drains many smaller
rivers along the way creating a huge watershed—the Amazon basin—which is
generally located below 200 m a.s.l. Associated with this watershed is an even larger
rainforest located in the equatorial regions of the South American countries of
Brazil, Columbia, Ecuador, Bolivia and Peru. Besides the Andes and the Atlantic
Ocean, the rainforest is bounded to the north by the Guiana crystalline shield and to
the south by the Brazilian crystalline shield (Pires & Prance, 1985) marked at their
edges by cataracts in the rivers.
This Amazonian rainforest is the world’s largest tropical rainforest and its largest
continuous forest of any kind, encompassing over 6,000,000 km2 (Holdridge, 1967;
Walter, 1973). It is also the most productive and diverse terrestrial ecosystem on
earth (Daly & Prance, 1989), containing more than 10% of all the species on the
planet (Pires & Prance, 1985). Not surprisingly this rainforest influences the entire
world’s precipitation and weather patterns as well as, over the longer term, the
world’s climate (Keller et al., 2004). Indeed, the rainforest may even control how
much rainfall it itself receives (Pires & Prance, 1985). Perhaps most importantly for
the human future the Amazonian rainforest interacts intimately with the Earths
carbon (C) cycle acting both as a carbon “sink”, by taking in large amounts of CO2
through photosynthesis, but also as a carbon “source” as, for example, when its
plants decay or burn (Rice et al., 2004). Further, the Amazon will continue to be a
major C player in the future by both contributing to, and suffering the effects of,
global warming (Shukla et al., 1990).
Within the Amazonian rainforest is the unflooded terra firme forest, located in
areas lower than 100 m elevation. The terra firme has the same physiogomy and
many of the same structural characteristics of similar unflooded rainforests
throughout the rest of the Neotropics (Kalliola et al., 1991; Lopez & Kursar, 1999;
Pitman et al., 2001; Myster & Santacruz, 2005; Myster, 2007a, b). For example the
Amazonian terra firme also contains a large amount of above-ground biomass and a
complex strata of emergent trees, canopy trees, understory trees, understory shrubs,
saplings, seedlings, herbs and ferns. Along with this vertical structure, it also has an
extensive horizontal structure of various-sized “gaps”, light flecks, micro-topological
relief and patchy soil nutrient availability in acid, clayey-loamy shallow soils with
extensive organic matter in its upper layers. As elsewhere in the Neotropics, many of
the terra firme trees have large buttresses and shallow roots, complex growth/
sprouting architectures, and large epiphyte and liana communities. Here, as throughout
the Neotropics, light availability is thought to be the most important environmental
factor controlling dynamics (Bazzaz & Pickett, 1980). However it may be the
interaction between light and water availability that is most important, simply because
when water is abundant plants grow large enough to shade out others making light
limiting, which they cannot do when water is limiting (Pires & Prance, 1985).
Flooding within the Amazon basin generates floodplains which cover approxi-
mately 3,000,000 km2, two thirds of which is “white” water (várzea: Junk, 1984)
Plant Communities of Western Amazonia 273

from the Andes. The water quality, frequency, duration, depth, and spatial variation of
this flooding gradient (Whittaker, 1975; Myster, 2001) greatly affects the distribution
and abundance of plant species (Junk, 1989; Lamotte, 1990; Ferreira & Stohlgren,
1999) leading to inundated vegetative associations created by the rise of the water table
on a regular, seasonal basis. In general forest-types, vegetation formations, and plant
communities run a continuum defined by the duration of the aquatic and terrestrial
phases of the annual cycles, and the physical stability of the habitat influenced by
sedimentation and erosion processes. In general the soils in the floodplain are less acid
than those of the terra firme but they have a greater concentration of exchangeable
cations such as Mg+2 and Na+2 (reviewed in Honorio, 2006).
The high species richness of the surrounding Amazonian terra firme rainforest—
when combined with this flooding and its associated environmental heterogeneity—
suggests that flooded forests will have a unique biology and ecology (Kalliola et al.,
1991). Indeed, it is expected that the regularity of flooding has lead to the evolution
of specific traits (Parolin et al., 2004) which then combine with flooding to create
specific tree species zonational distributions, largely determined by the submergence
tolerance of their seedlings.
The section of the Amazonian rainforest covered in this review—Western
Amazonia—is here delineated west from Manaus, Brazil (@ 60° W longitude) at
the confluence of the Amazon and Negro rivers (Fig. 1). This part of the Amazon
basin is composed of young and relatively fertile sediment deposited subsequent to
the initiation of the Andean orogeny in the Miocene making the flora of more recent
origin than that of eastern Amazonia (reviewed in Dumont et al., 1990; Terborgh &
Andresen, 1998). The western Amazon rainforest may be of particular scientific and
social interest because it contains some of the most diverse areas or “hotspots” on
earth (Myers et al., 2000; Myster, 2007a, b). Indeed, while the Amazon in general
suffers from human habitation and behaviors—for example 40% of it may be farmed
by the year 2050—the western Amazon remains largely intact, pristine, and
unaffected by humans with less than 5% of it’s total area in secondary forest (Neff
et al., 2006; Soares-Filho et al., 2006).
This review is motivated by the need to define and refine community classifications
in the Western Amazon beyond just the unflooded terra firme and a couple of basic
flooded forest-types (Terborgh & Andresen, 1998). Towards that end, this review uses
data from conducted vegetation surveys and studies, discussions from book chapters
that focus on Amazon plant communities (e.g., Pires & Prance, 1985), and my own
observations when working at three different field stations in the western Amazon.
Further because the Amazon rainforest has at least 30,000 plant species, because its
beta diversity is not low (Gentry, 1988) and because it has extensive variation in
edaphic, climatic, topographic and geological conditions, a large number of distinct
plant communities are expected. Indeed, sampling has shown that up to 59% of
recorded species are found in only one plot among many sampled (Duivenvoorden,
1995) and shared species between wetland and upland forests is very low (Dumont et
al., 1990). As you shall see studies suggest that the Amazonian rainforest, and in
particular the Western Amazon, itself may contain more different species of plants, and
different kinds of plant communities, than the rest of the American hemisphere put
together! If so that would make the Western Amazon one of the “last frontiers” for
plant scientists of all kinds.
274 R.W. Myster

Fig. 1 Map of the 19 major tributaries of the Amazon river. These rivers in the Western Amazon are
white-water (Japurá, Napo, Marañon, Ucayali, Juruá, Purus, Madeira, Icá, Javari) or black/clear-water
(Negro, Branco, Jutai, Guaporé). Places where these rivers and/or their floodplains meet would be mixed-
water areas (reprinted with permission from World Fisheries Trust)

Western Amazon Plant Communities

Unflooded

Terra Firme Low Terrace

The most common forest-type within the terra firme may be this kind of unflooded
forest not located in ridges next to rivers. It forms in clay soils which are higher in
Nitrogen (both ammonium and nitrate) than white-sand forests, and its soil
characteristics may be locally important, for example Tuomisto et al. (2002) found
a negative correlation between soil extractable bases and both plant density and
species richness, and Ruokolained et al. (2007) report that soil differences explained
Plant Communities of Western Amazonia 275

50% of floristic differences among terra firme plots while geographic distances
explained only 16%. This forest-type has a maximum height of at least 20 m and its
total above-ground biomass has been estimated as 397.7 Mg/ha (including all woody
plants plus palms: Nascimento & Laurance, 2002) and 405–560 t/ha (Ferreira &
Prance, 1998). Decomposition studies have yet to be done here, but (I suggest)
should include both the overall rate of decay and mineralization (see Myster &
Schaefer, 2003).
Sampling for all woody plants and palms >1 cm diameter at breast height (dbh) in
a 25 ha plot found 1,104 species, 152,353 individuals, and the dominant families
Bombacaceae, Palmae, Lecythidaceae, Fabaceae and Violaceae, and in a 0.2 ha plot
found 413 species/ha (estimated) with 96 liana species/0.2 ha (reviewed in Leimbeck
& Balslev, 2001). Sampling all trees less than or equal to 2.5 cm dbh found an over
abundance of ferns and trees in the family Melastomataceae (Duivenvoorden et al.,
2001). Sampling for trees >5 cm dbh showed 473 species/ha and 1,561 (1,509 in
Leimback & Balslev, 2001) trees/ha (Valencia et al., 1994, 2004).
Several studies have sampled trees >10 cm dbh showing: (1) for density 693
individuals/ha (Valencia et al., 1994, 2004), 649/ha (Boom, 1986), 639–713/ha
(Ferreita & Prance, 1998), 734/ha (Korning et al., 1990), 728/ha (Balslev et al.,
1987), 524/ha (Spichiger et al., 1996) and 602/ha (Campbell, 1994), (2) for species
307/ha (Valencia et al., 1994, 2004), 94/ha, 137–168/ha (Ferreita & Prance, 1998),
182/ha (Campbell, 1994), 228/ha (Balslev et al., 1987), 153/ha (Korning et al.,
1990), 227/ha (Spichiger et al., 1996) and 65–285/ha (reviewed in Worbes, 1997),
and (3) for basal area 33.7 m2/ha (Balslev et al., 1987), 22.2 m2/ha (Korning et al.,
1990), 23.6 m2/ha (Spichiger et al., 1996), 21.4 m2/ha (Boom, 1986), and 32.8–
40.2 m2/ha (Ferreira & Prance, 1998). Common families in these studies were
Leguminosae, Palmae, Moraceae, Sapotaceae, Lauraceae, Lecythidaceae, Chrysoba-
lanaceae, Myristicaceae and Melastomataceae (Boom, 1986; Spichiger et al., 1996,
Terborgh & Andersen, 1998; Ferreira & Prance, 1998; Honorio, 2006). However a
forest sampled in Southern Venezuela that seemed to be terra firme had only
83 tree species/ha and only 84% of the basal area of a similar plot in Mananas (at
dbh >10 cm: Uhl & Murphy, 1981). Such variation among 1 ha plots shows that
more sampling needs to be done before categorizing the terra firme low terrance
forest.
The few forest studies of plant-plant replacement mechanisms (key to forest
dynamics: Myster, 2007a) have included both seed predation (Romo et al., 2004;
Wyatt & Silman, 2004), herbivory, and seedling survivorship (Fine et al., 2006). The
herbivory study showed that plants lost 23% of new leaves per month to herbivores
compared to only 10% per month in white-sand forest, however there was more
seedling mortality in white-sand forest due to total defoliation (Fine et al., 2006).

Terra Firme High Terrace

This forest is located next to rivers (called barrancos for the terra firme forest that goes
right to the high river bank: Pires & Prance, 1985) but is rarely, if ever, flooded. It has
20–30 m tall trees with a higher diversity of palms than floodplain forest, but a lower
diversity compared to palm swamp communities in terra firme. One sampling found
15,869 individual plants and 37 species of palms in 2.5 ha with the most common
276 R.W. Myster

species Phytelephas tenuicaulis, Astrocartum murumuru, Attalea butyracea, Iriartea


deltoidea, Aphandra natalia and Hyospathe elegans (Normand et al., 2006).

Palm Swamps

These swamps (called bosques de quebrada: Kahn & Mejia, 1990), occur in
depressions or low-lying areas of poor drainage within terra firme. They get their
water from rainfall and their soil is waterlogged for a longer period of time than the
floodplains. They are known for low tree diversity, dominance by a few species, and
a marked species similarity between countries (Vormisto et al., 2004). The forest
forms a canopy layer 12–30 m in height with crowns of tree-like palms intermixed
with larger trees. Whereas common families include Arecaceae, Clusiaceae,
Fabaceae, Lecythidaceae, and Humiriaceae (Honorio, 2006), monocots are 12–
31% of the species. In particular palms can form pure, dense stands (e.g. Mauritia
flexuosa or “morichal”) with species such as Iriartea deltoidea and Geonoma
macrostachys common (Svenning, 1999). The local distribution of palms and
species with a palm-like habit are strongly influenced by topography and drainage.

White Sand

These forests (called caatinga) form on white sand soils of extreme infertility, made up
mainly of quartz podzols. As in swamps, but unlike in terra firme (Duivenvoorden &
Lips, 1995), single species tend to dominate (e.g., Micrandra sprucei, Eperua
leucantha) with individual trees having slender boles and roots mainly at the soil
surface (Klinge et al., 1990). This forest has been divided into varillal (dense with
straight and thin trees between 10 and 20 m high) and chamizal (shrubs 3 m high and
scattered trees <8 m high: Honorio, 2006). There may also be a formation of heath-
like vegetation (bana) with low productivity, low diversity, low canopy height and a
small amount of fine root biomass (only 5.5 kg/m2/ha compared to 10.75 kg/m2/ha in
caatinga and terra firme: Sanford, 1989). Likewise its species diversity (for trees with
a dbh >10 cm) is between terra firme (which is higher) and palm swamps (which is
lower: Duivenvoorden, 1996). Common families include Leguminosae, Clusiaceae,
Malvaceae, Euphorbiaceae and Icacinaceae (Honorio, 2006).

Savanna

Found on sandstone plateaus and outcrops between 300 and 350 m a.s.l elevation and
having few trees, this kind of formation is mainly made up of shrubs, herbs, lichens,
liverworts, and mosses (called campos rupestres locally). Some savannahs may have no
woody vegetation and/or be inundated (Pires & Prance, 1985). Further subdivision of this
forest-type into caatinga savannas found on loose sediments (Duivenvoorden & Lips,
1995) and campina savannas on white sand (Pires & Prance, 1985) has been suggested.

Gaps/Forest Floor Heterogeniety

Researchers have historically regarded the rainforest as a “mosaic” composed of


areas of closed canopy and areas of past and present treefall gaps in various stages of
Plant Communities of Western Amazonia 277

regeneration (Watt, 1947) and it has been shown for terra firme seedlings that they
grow more in gaps, where both light and water availability are higher, compared to
the understory (Poorter & Hayashida-oliver, 2000). My own sampling in terra firme,
high restinga and tahuampa forests and their gaps found there were fewer, but taller,
bigger, and more productive trees in the forest plots compared to the gap plots and
that although the most flooded plots had the fewest trees and species, they also had
the greatest amount of turnover. Interestingly the intermediately flooded forest had
the greatest richness and alpha diversity supporting a “mass effects” hypothesis
where species from both the unflooded and most flooded forest and their gaps have
overlapping ranges in the medium flooded forest and gap (Myster, 2007b).
Further experiments in forests and gaps showed that (1) the total number of
dispersed seeds was greatest in the high restinga gaps with more seeds falling in
the gaps for both terra firme and high restinga and greater seed richness in both
gaps compared to their forest, (2) the greatest number of seedlings germinated from
high restinga gap soils with the least from the terra firme forest, and gaps had more
seedlings germinating from their soil samples compared to their respective forests,
the majority of seeds were lost to predators (only approximately 30% remained
after two weeks in the field) and species were significantly different in most
situations, (3) large seeds suffered more losses to pathogens than small seeds and
pathogenic losses were greater in the forests compared to their gaps, more seeds
germinated in the wet forest compared to the dry forest, and (4) seedling species
differences dominated over site differences, as seed mass increased seedlings
survival rates increased but growth rates declined, and more seedlings survived in
gaps compared to closed forests where they also grew faster with a larger leaf area
ratio (author, unpub. data).
However the utility of this forest/gap dichotomy has been increasingly called into
question (Lieberman & Lieberman, 1989) because of the large environmental
variation within gaps, because the boundary of forest and gap has been found to a
gradient from light to dark and not a distinct “hole” in the canopy, and because the
forest itself contains persistent openings of light (Chazdon & Pearcy, 1991). This
suggests that forest floor sampling should be done and, indeed, sampling done for
herbaceous plants on the forest floor of the terra firme low terrance forest showed all
were perennials with 25 species of ferns and 71 species of angiosperms per ha, the
most common families were Araceae, Marantaceae and Poaceae, and approximately
2.5% of the forest floor was covered in herbs (10,960 individual plants/ha) of which
41% of the species had a climbing or epiphytic habits (Poulsen & Balslev, 1991).
My own sampling along a long transect in Amazon terra firme showed that (1)
the spatial heterogeneity across the transect was greater than the temporal
heterogeneity at any given point on the transect for both light and water and that
there was a positive correlation between light availability and soil water, (2) Couepia
obovata, the largest seeded and the only sub-canopy tree, survived best and showed
both the largest relative growth rate (RGR) and the largest specific leaf mass (SLM)
while for the two trees, both early successional, Tapirira guianensis had the largest
leaf area ratio (LAR) and the largest leaf mass ratio (LMR) and Duguethia spixiana
had the largest root/shoot ratio, (3) for T. guianensis while light predicted increasing
SLM with increasing light, soil water potential predicted both increasing LMR and
decreasing root/shoot ratio with increasing soil water, and (4) soil water potential
278 R.W. Myster

could also predict increasing LAR with increasing water for D. spixiana and, for C.
obovata, soil water potential predicted more survivorship, LMR, and RGR but less
root/shoot ratio all with increasing soil water (author, unpub. data).

Post-Agricultural/Pasture

The conversion of rainforest to agriculture and pasture is the most common


disturbance in the Neotropics human or natural (Myster, 2007a) where cleared areas
are abandoned within a few years of cropping due to thin soil and substantive
leaching and loss of nutrients. This leads to loss of productivity, for example NPP
can drop to only 7–9 t/ha after 12 years of cropping. The favorite of traditional
agriculture, Cassava, is commonly done in terra firme, but not in flooded areas
because it rots easily. Also the high aluminum content in the soil generally means
that banana, pineapple and manioc do well.
In addition to soil nutrients, species richness also declines with both time and
extended use—235 (forest)=>180 (cropped field)=>150 (fallows)=>140 (cropped
again)=>81 (pasture)—with trees, palms and vines most common in forest, shrubs
most common in the middle of the sequence, and herbs most common at end
(Fujisaka et al., 1998). Chronosequences suggest that species richness will recover
back to forest levels after 20–40 years post-abandonment, and that after 10 years
basal area will be 11 m2/ha compared to 37 m2/ha in mature forest and aboveground
biomass will be 44 vs. 326 t/ha for mature forest (Saldarriaga et al., 1988).

Post-Logging/Wildfires

Two common logging techniques in the Amazon are tree-thinning and strip clear-
cutting. One study showed that while tree thinning enhanced growth of the remaining
trees, it had no better effect than strip clear-cutting on recruitment, survival and species
richness (Dolanc et al., 2003). Regeneration studies showed that seed rain (mainly from
birds) declined exponentially from forest to edge of logging strips (i.e., ~1,000 seeds/
year/16 m2) with the most common families Melastomataceae and Moraceae (Gorchov
et al., 1993). However, mammal-dispersed species have trouble dispersing into logged
areas (e.g., Hymenaea courbaril: Gorchov et al., 2004). For dispersed seeds, predation
showed a loss rate of 30–80% removed after 25 days depending on species and
microsite, where the edge between forest and logging strip had the greatest losses
(Notman et al., 1996).

Other

Several other unflooded forest types have been observed and commented on in the
literature. For example, a separate liana forest may be possible which forms in rich
mineral deposits on slightly higher area (Pires & Prance, 1985). If so, then such a
forest may become more common in the future because lianas are increasing in
undisturbed terra firme forests, alms with increasing tree mortality and suppressed
tree growth (Phillips et al., 2002). Also a “pre-Columbian” forest may exist, a result
of old human settlements (Campbell, 1994). In addition, authors have commented on
a dry, low forest made of open sand clumps of bushes and small trees (campina:
Plant Communities of Western Amazonia 279

Guillamet, 1987) and a campinaranalow forest of thin-stemmed trees, 10–20 m high


with large broad-trunked individuals (Guillament, 1987). There may also be an
“open” forest both with and without palms (Pires & Prance, 1985). Finally there may
be a creek-side forest with 81 palms/ha and arborescent palms (7% of the total:
Peres, 1994) and an open forest, with less biomass, smaller trees, shrubs and lianas,
with fewer epiphytes. The reviewed forest types (especially savannahs) and many of
those in this other category can be found in Mayle et al. (2007).

Seasonally Flooded with “White” Water (Várzea)

Floodplains are flat, well-drained areas above mean river levels. Within the Amazon
basin, rivers are either “white-water” consisting of runoff from the Andes or “black-
water” derived from local catchments of rainfall moving through forest litter. Both kinds
of rivers help define dynamic areas where soil erosion and sediment deposition change
patches quickly. White-water rivers are muddy, rich in sediment and suspended solids,
high in pH (6–7) because they derive from sedimentary deposits, and high in dissolved
inorganic substances with a high percentage of cations (P. K, Ca) and alkali-earth metals
compared to alkali metals. The soils they create by deposition are very fertile which may
help to offset the general low fertility of Amazon soils.
Várzea is the general name given to the large fringing floodplain of white-water
in the Amazon. Sampling has shown that the várzea is more productive (33.6 t/ha/
year for the woody vegetation and 50–100 t for the herbaceous vegetation) than
terra firme (12–27 t: Worbes, 1997), with the productivity of 12-year-old Cecropia
latiloba stands alone at 8.1 t/ha/year (Parolin, 2002b). One sampling had 40 species/
ha (Worbes, 1997) which, although smaller than terra firme, has more individuals
per species. Within várzea, species may increase as you move west or with a
decrease in flood duration.
Because flooding is so regular here, plants in várzea time much of their life-cycle
around flooding, producing a unique ecology. Although the start of the flooding cycle
produces an increase in leaf fall, and a reduction in both growth and leaf N (Parolin et
al., 2002b), flooding also leads to peak fruiting and seed dispersal into the water, where
seeds float, and are often eaten by fish. If a seed is protected by a thick seed coat or are
very small (Junk, 1989), it can pass through a fish and germinate when floodwaters
recede (Kubitzki & Ziburski, 1994) or suffer secondary predation. Specific studies
have shown a seed predation loss rate of 8%/week for Dipteryx micrantha
(Leguminosae) and palm seedling survivorship negatively correlated with level of
flooding (Losos, 1995). Studies have shown that tree seeds that are under water less,
germinated earlier and the resulting seedlings were taller with longer leaves (Parolin et
al., 2003) than those seeds that are underwater more. Another study showed that
Himatanthus sucuuba germination was above 80% for seeds that were under water
even up to 120 days, and it’s seedlings responded morphologically under root
submersion, and physiologically (higher values of the enzyme alcohol dehydrogenase)
under total submersion, with 70% survival after 120 days (Ferreira et al., 2006).
Flooding reduces leaf growth and total biomass for almost all tree species (Lopez
& Kursar, 2003). After germination seedlings must grow fast so that leaves are
above water level the next time the river floods. Alternatively, seedlings must
develop the ability to endure extended submersion and many species of seedlings
280 R.W. Myster

can tolerate submergence for a few weeks, with a new leaf flush once waters recede.
There are at least two general strategies for seedlings in floodplains (1) grow fast with
new seedlings having high shoot elongation or (2) tolerate submergence with an array of
specific morphological and physiological traits (Parolin, 2002a, b, c, d). In addition
resprouting also occurs when rivers first recede often due to rotting and mechanical
damage. Although most growth occurs in the unflooded season, various anatomical and
morphological adaptations allow photosynthesis rates not to decrease during flooding
(Parolin et al., 2004), for example aerenchyma root tissue to facilitate gas exchange
under water (Junk, 1989; Lopez & Kursar, 1999). Other important morphological
adaptations include hypertrophy of lenticels, adventitious roots, plank-buttressing, stilt
rooting, aerenchyma, suberin and ligin in the root peripheral cell layers.

High Restinga

Located between 25 and 28 m a.s.l, this forest (várzea alta) is flooded 1 month/year
with a water level of 1–2.5 m (Wittman et al., 2002), which is low enough for many
terra firme tree species to establish (Wittman et al., 2004). Also this forest has larger
trees than terra firme forests and a faster tree turnover (23 vs. 37 years for nearby
terra firme: Korning & Balslev, 1994). One sampling found that leaf litter fall in this
forest was 700 g/m2, above-ground biomass was 34–48 kg/m2, and net primary
productivity was 2,082–2,558 g/m2 (Nebel et al., 2001c).
Other sampling has shown this forest has (1) a density of 1,012 trees/ha (dbh >
5 cm: Leimbeck & Balslev, 2001;Worbes, 1997; Klinge et al., 1990), or 417/ha (dbh >
10 cm: Balslev et al., 1987) and 500 trees/ha (dbh > 10 cm: Worbes et al., 1992) for
larger trees, (2) a species richness of 236 species/ha (dbh > 5 cm: Worbes, 1997;
Klinge et al., 1990), or 149 species/ha (dbh > 10 cm: Balslev et al., 1987) and 88
species/ha (Worbes et al., 1992: dbh > 10 cm) for larger trees, and (3) a total basal area
of 41–60 m2/ha (dbh > 5 cm: Worbes, 1997; Klinge et al., 1990), or 35.5 m2/ha (dbh >
10 cm: Balslev et al., 1987) and 4.8 m2/ha (dbh > 10 cm: Worbes et al., 1992) for
larger trees. When species richness is normalized to 100 stems, there were 9–18
species/100 stems (Worbes et al., 1992). If shrubs are sampled with trees >10 dbh,
there are 4,458 individuals/ha with 208 species in the overstory and 204 in the
understory (Nebel et al., 2001a). The most common families include Myrtaceae,
Fabaceae, Euphorbiaceae, Arecaceae, Lecythidaceae, Rubiaceae and Moraceae
(Godoy et al., 1999; Honorio, 2006).

Low Restinga

Located 21–25 m a.s.l, this forest-type (várzea baixa) floods 4–6 months/year with a
water level between 2.5 and 5 m and sampling has shown 2,330 individual trees/ha,
35 species/ha and 2.6 m2 total basal area/ha (dbh > 10 cm: Wittmann & Junk, 2002).

Chavascal

This forest-type floods 6–8 months/year with a water level of 5–7 m (Wittmann et
al., 2004).
Plant Communities of Western Amazonia 281

Swamp

There is only one report of a permanent whitewater swamp forest (Prance, 1979) but
without any sampling.

Gaps/Forest Floor Heterogeniety

Regeneration in this forest may happen only in certain years when gaps occur after
the water has receded and uprooting can happen in the wet soil. Although not
studied, the expectation is that the resulting gap dynamics and regeneration in
flooded areas should also have unusual aspects (Lamotte, 1990).

Post-Agricultural/Pasture

The fertile soils here and the proximity to transportation make these areas
attractive for farming (Padoch et al., 1999), but the logistics make it difficult.
However it may be possible using flood-resistant crops, precise timing and careful
zonation away from the most active regions. If farming is tried, mud beaches are
prime areas especially for rice and vegetables, and native plants, esp. palms, may
be harvested for food (e.g., Euterpe oleracea) and fibers (e.g., Mauritia flexuosa).

Post-Logging/Wildfires

These areas are attractive for logging because of easy transportation but the
flooding must be infrequent and the water level low. Selective logging, of
Virola (Virola surinamensis) for example, may be most possible (Padoch et al.,
1999). One sampling suggests that heavy felling can lead to more diameter growth
in the remaining trees (6.8 mm/year compared to 4.5 elsewhere: Nebel et al.,
2001b).

Other

Over time, rivers form new floodplains and abandon old ones. These old
floodplains contain abandoned river channels, depressions, oxbows lakes, and
insular sedimentary beds of different ages, all of which may form their own unique
vegetation. These meandering rivers thus create vegetative mosaics (Watt, 1947)
related to differences in water and soil chemistry, mode of alluvial sedimentation,
and bio-geographical events (Salo et al., 1986). Channel patterns include
meandering, braided and anastomosing, which may define different successional
communities (Puhakka et al., 1992). Parodi and Freitas (1990) also speak of
floating meadows, riverine forest on old beaches, and creek forest, and other
community types may include fluvial bars, swales (bajeales grasses leading to two
shrubby stages of succession: Lamotte, 1990), abandoned channels, and riverbanks
which form a successional sequence (Kalliola et al., 1991). Also pure stands of
palm are seen in the middle of white-water rivers (Buritizal: Pires & Prance,
1985). In particular, regeneration on newly exposed fluvial deposits (beaches) may
282 R.W. Myster

be an example of primary succession (see Myster, 2001 for other examples).


There may be a distinction between clay vs. sand várzea, with clay várzea having
higher river banks (Pires & Prance, 1985). Finally there may also be a nonforested
area of shrubs, herbs, and ferns located next to white-water rivers (campos de
varzea: Pires & Prance, 1985) and additional forest formations (Kvist & Nebel,
2001).

Seasonally Flooded with “Black” Water (Igapó)

These rivers originate in the Amazon basin or its adjacent crystalline shields and
have a low amount of sediments. However they are rich in tannic and humic acid,
and degradable phenolic substances (organic matter) because they are decomposition
runoff from forest litter. Such processes usually produce a broad-leaf forest (Parodi
& Freitas, 1990) with zonation differences for different plant species, but they are
not as distinct as in white-water (Ferreira, 2000).

Tahuampa

Perhaps because plant zonation is less pronounced here compared to várzea forests,
only one general type of forest is related in the literature (Prance, 1979). Its forest
trees have a tortuous growth form with boles grown together, bunches of
adventitious ftious roots, and a bark that may be able to withstand fire and repair
themselves afterwards (Pires & Prance, 1985).
This forest has a low diversity of palms with (1) a density of 222 trees/ha (Klinge
et al., 1990) and 502 trees/ha (Campbell, 1994) for stems with a dbh >10 cm but
1,880 trees/ha for trees of all sizes (Parolin et al., 2003), and (2) a richness of 96
species/ha (similar to várzea: Campbell, 1994) and 80 species/ha (Worbes, 1997) for
trees with a dbh > 10 cm (Klinge et al., 1990). Most common families included
Leguminosae, Annonaceae, Palmae, Moraceae, Sapotaceae, Euphorbiaceae and
Lecythidaceae (Honorio, 2006).

Palm Swamp

Permanent black water swamp forest is mentioned in Parodi and Freitas (1990) and
Prance (1979) with common families Palme, Arecaceae, Euphorbiaceae, Fabaceae
(Duivenvoorden & Lips, 1995; Honorio, 2006).

Post-Agricultural/Pasture

The low fertility of soils make agriculture unlikely here.

Post-Logging

Selective logging may occur here but it is not mentioned in the literature.
Plant Communities of Western Amazonia 283

Other

As for white-water plant communities, meandering rivers produce point bars or


beaches after receding and repeated sedimentation can create ridges separated by
depressions called swales (Kvist & Nebel, 2001) creating flooded woody (bajial,
pungal, aguajal) and non-woody (pantanal, barrial, playa) associations. Sampling
has shown for trees with a dbh > 10 cm, these densities for a lake (777 trees/ha),
river margin (941) and stream (1,111), and these species diversities for a lake (44
species/ha), river margin (103 species/ha) and stream (137 species/ha) with common
families Leguminosae and Euphorbiaceae and total basal area for a lake (22.3 m2/
ha), river margin (41.8 m2/ha) and stream (34.2 m2: Ferreira, 2000).
There are also grassland communities that flood seasonally with black-water
(Haase & Beck, 1989). These include both herbaceous-dominated plant
communities—a “dry” bunchgrass savanna, a seasonally flooded bunchgrass
community, a palm swamp whose understory is dominated by grasses and sedges,
a floodplain community similar to those previously described but with a greater
abundance of grasses, a river bank community—shrub communities, a community
arising from termite mounds, and a tall gallery forest (Haase & Beck, 1989).
Finally rivers may also create white-sand islands during the dry season (Pires &
Prance, 1985).

Seasonally Flooded with “White” and “Black” Water

One sampling of this forest showed that for trees with a dbh >10 cm, there were
396 individuals/ha, 132 species/ha, and 30 m2 total basal area/ha. This included
species from both white-water and black-water families (e.g, Euphorbiaceae,
Sapota, Clusiaac, Mimosa, Legumes: Amaral et al., 1997), which suggests that
vegetative formations here consist of a mix of species from other forests rather than
a completely new group of species (Encarnacion, 1993). These potential
communities are also mentioned by Prance (1979). Further complexity is implied
when clear water may flow into white-water rivers during low flow times and
white-water may flow into black water rivers at the beginning of the rainy season
(Puhakka et al., 1992).

Aseasonally or Irregularly Flooded

These forests are flash-flooded by irregular rainfall, and exist mainly in the upper
reaches of rivers and beside small creeks (Prance, 1979). Flooding may happen at
any time of year, but is usually of short duration so that some terra firme species are
found here (Prance, 1979). One sampling has shown 417 trees/ha, 149 species/ha
and 35.5 m2 total basal area/ha (Klinge et al., 1990). Located at 285 m a.s.l, these
communities can include permanently flooded swampy depressions (aguajal)
located parallel to river beds and getting their water from irregular flooding (Kahn
& Mejia, 1990). These sites may be higher in species than the seasonal or
periodically forested forests (Dumont et al., 1990).
284 R.W. Myster

Synthesis

Information currently known and sampled about plant communities in the Western
Amazon may begin to answer these basic questions:

1. Are there trends of forest-type that correspond to flooding gradient parameters


such as maximum level and duration of flooding?
One comparative sampling found that for trees of dbh < 10 cm, terra firme had >
220 species/ha, flooded forests had < 150, palm forest had 60 and white sand forest
had < 100/ha (Honorio, 2006). Another one showed 196–249 species per 0.1 ha in
terra firme vs. 163–168 in flooded forests (Gentry, 1986). It seems that as an area
becomes more flooded with white-water the number of trees increase and the
number of species decreases (Campbell et al., 1992), i.e. terra firme is richer than
várzea which is richer than igapó, with species number decreasing with increasing
flood-period everywhere. In addition floodplain trees are shorter than trees in terra
firme, but basal area in várzea higher (Klinge et al., 1990). There are more epiphytes
in white-water forest than in terra firme (Leimbeck & Balslev, 2001) and igapó has
more evergreen leaves than várzea.
2. Are white-water and black-water communities different, when standardized for
water level and flooding duration?
It is known that várzea has higher productivity with faster tree growth and lower
wood densities than igapó, and that fast-growing “pioneer” trees are absent in igapó
with genera Cecropia, Salix, and Senna common in várzea (Parolin, 2000a, b).
3. Are riverbank communities successional (Wittmann et al., 2004)?
There is a sequential appearance of species along topographic gradients going
away from rivers due to levels and ages of sedimentation and silting, and succession
may take place on the convex banks/point bars (Lamotte, 1990) with positive
feedback between trees and sediments. If one accepts this idea that flooding
communities are to a degree “successional” then one parallel is that early
successional trees (e.g., Cecropia latiloba, Saliz humboldtiana, Senna reticulata:
Parolin et al., 2002b) establish on areas that are often flooded, having expected traits
of high germination success, light wood and fast growth rates. Cecropia latiloba in
particular establishes very well on white-water river sides because it can tolerance
waterlogging (lack of O2), and sedimentation, and has fast vertical growth (21 umol
CO2/m2/s; Parolin, 2002b).
4. Are there trends in regeneration across forest-types?
There is a higher seed mass in igapó compared to várzea and within the flooding
gradient in igapó seeds are heavier if less flooded, but in várzea no pattern. I found
that the greatest seed and seedlings losses—and the slowest seedling growth rates—
were in várzea gaps with medium losses and growth rates in terra firme gaps and
várzea forest, and smallest losses and growth rate in terra firma forest (author,
unpub. data). Other studies have shown lower and later germination rates trees from
várzea compared to igapó (Parolin, 2001a, b). Interestingly assimulation rates of
flooded trees, when they are not flooded, is the same as terra firme trees (Parolin et
al., 2001). In another study I found that while canopy structure is determined by
traditional gap dynamics, much of canopy diversity depends on the forest type,
density decreases as flooding increased, especially among the smallest stems, and
Plant Communities of Western Amazonia 285

the high biodiversity of these areas may be maintained in part by the existence of
moderately flooded forest and gaps, suggesting how they may be conserved (author,
unpub. data).
5. Are there clear-water vegetation types that are not found in black-water?
The distinction between “clear” water and black-water is not made by many
authors and actually sampling is rare. Whether this distinction is real or only
apparent remains an open question. However, even if the water quality is
different, the real question is whether plants respond differently to it and if they
then form distinct plant communities. In general vegetation differences have not
been explored.
Taken together studies suggest that the traditional “pioneer” tree strategy may be
too simple for flooded forests, and a “fugitive” strategy (height growth without
branching, and tolerance to submergence) and a “possessive” strategy (growth with
lateral branching and shading of competitors: Parolin, 2003) are needed. Drought
and waterlogging produced a similar amount of stress and similar growth responses
in seedlings afterwards (Parolin, 2001b), and so it may be that some sub-canopy
trees may survive and grow more than early-successional or “pioneer” trees when
presented with water stress in the forest understory, and that within the normal
ranges of light and soil water plants responded more to spatial variation in soil water
availability than in light, contrary to the traditional view that light is the most
limiting resource in tropical forests (Bazzaz & Pickett, 1980).

Conclusions

We have seen evidence for a large variety of plant communities in the Western
Amazon. Most of it was sampled data from plots or transects, some of it was the
product of the historical experience of the local people who live there, but some was
only educated observations and so can best be described as anecdotal. For example,
Terborgh and Anderesen (1998) showed that flooded forests tend to have species in
common more with close-by terra firme forests than with flooded forests farther
away and that Amazon plant communities fall out on soil fertility gradients and
length of dry season. This was supported by Pitman et al. (1999) who reports that
most plant species in western Amazonian are wide-spread, usually with a density of
< 1 individual per ha, with only approximately 15% species shared between forest-
types (Pitman et al., 1999).
Taken together studies suggest that when the vegetation of the Western Amazon is
sampled as completely as plant communities have been in temperate America, we
will see a very large number of distinct communities. This is further suggested from
the large number of plant species in the Amazon and the effects of environmental
heterogeneity (e.g., soil cation content: Tuomisto et al., 2003) seen, for example,
when species partition flooding levels into vegetative zones.
Even so, the definition of a plant community and the best characteristics and
methodology to use for its quantification remains an open question (e.g., assembly
rules, discreteness, integrateness: Wilson 1991). Computational attempts to define
communities in the Western Amazon have generally used multivariate ordination
286 R.W. Myster

(Duivenvoorden & Lips, 1995) and clustering algorithms on similarity indexes


between plots to define them as “in” or “out” of a community-type. Such methods,
for example, did define six plant communities in the Colombian (Duivenvoorden &
Lips, 1995) and Peruvian Amazon (Honorio, 2006). But such kinds of analysis
decide beforehand how many communities there are and then merely examine how
well plots fit into them, without a statistical significant test of association either
between plots or between species.
I would like to propose that such investigations begin with significant
associations between species and then use those to build species groups and plant
communities hierarchically (Holling, 1992). I have done this with post-agricultural
plot data from Ecuador (Myster, 2007a; author, unpub. data) starting with (1)
spearman rank correlation coefficients between all species pairs, (2) next building a
matrix where distances between species reflect their degree of positive association,
and finally (3) using that association matrix to cluster species into groups. Such an
approach eliminates the need for a “in or out” subjective placement as the
methodology for defining plant communities instead seeing them as existing on a
continuum—here a clustered hierarchical metric of association—just as their plant
species distributions are expressed individually on spatial and temporal gradients
(Whittaker, 1975; Myster & Pickett, 1988) and together on hierarchical gradients of
α, β, and γ species diversity (Whittaker, 1965).
Finally although some taxonomic descriptions of many Western Amazon plant
communities have been made, there has only a cursory description of basic structural—
gross, net, above-ground and below-ground productivity and gas exchange: Nascimento
and Laurance, 2002)—and functional—decomposition decay constant k and mineral-
ization rates (Myster & Schaefer, 2003)—parameters. Much work needs to be done. I
would hope that governmental agencies—for example the US National Science
Foundation—would see the importance of such work and reflect it in their funding
priorities. A series of LTER-style (Long-Term Ecological Research: see Myster,
2007a) research stations clustered around the plant community types reviewed here,
may be needed to fully understand the Amazon and monitor its functional changes in
the decades ahead, which is so important to the health of the Earth.

Acknowledgements I would like to thank Dr. Hugo Naverette, Dr. Renato Valencia, and the staff of the
Yasuni field station (Ecuador), Ms. Euidice Horonico and the staff at the Jenaro Herrera Field Station
(Peru), and Dr. Paul Beaver and the staff of the Tamshiyacu-Tahuayo Reserve (Peru) for their help in
facilitating my research in the Western Amazon. I would also like to thank Henrik Balslev, Evan Notman
and Michael Gilmore for commenting on a previous version of the manuscript. This research was
supported by NSF grants DEB-0218039 and DEB-0620910 to the Institute for Tropical Ecosystem Studies/
University of Puerto Rico and to the International Institute of Tropical Forestry as part of the long-term
ecological research program in the Luquillo Experimental Forest. Additional support was provided by the
Forest Service (U.S. Department of Agriculture) and the University of Puerto Rico.

Literature Cited

Amaral, I. L. D., J. Adis & G. T. Prance. 1997. On the vegetation of a seasonal mixedwater inundation
forest near Manaus, Brazilian Amazonia. Amazoniana 14: 335–347.
Balslev, H., J. Luteyn, B. Ollgaard & L. B. Holm-Nielsen. 1987. Composition and structure of adjacent
unflooded and floodplain forest in Amazonian Ecuador. Opera Botanica 92: 37–57.
Plant Communities of Western Amazonia 287

Bazzaz, F. A. & S. T. A. Pickett. 1980. Physiological ecology of tropical succession: a comparative


review. Annual Review of Ecology and Systematics 11: 287–310.
Boom, B. M. 1986. A forest inventory in Amazonian Bolivia. Biotropica 18: 287–294.
Campbell, D. G., J. L. Stone & A. Rosas. 1992. A comparison of the phytosociology and dynamics of
three floodplain (varzea) forests of known ages, Rio Jurua Brazilian Amazon. Botanical journal of the
Linnean Society 108: 213–237.
———. 1994. Scale and patterns of community structure in Amazonian forests. Pp 179–197 in P. J.
Edwards, R. M. May, & W. R. Wedd (eds) Large-scale ecology and conservation biology. Blackwell
Scientific, Oxford.
Chazdon, R. L. & R. W. Pearcy. 1991. The importance of sunflecks for forest understory plants.
Bioscience 41: 760-766.
Daly, D. G. & G. T. Prance. 1989. Brazillian Amazon. Pp 401-426 in D. G. Campbell, & H. D.
Hammond (eds), Floristic inventory of tropical countries. New York Botanical Garden, Bronx.
Dolanc, C. R., D. L. Gorchov & F. Cornejo. 2003. The effects of silvicultural thinning on trees
regenerating in strip clear-cuts in the Peruvian Amazon. Forest Ecology and Management 182: 103-
116.
Duivenvoorden, J. F. 1995. Tree species composition and rain forest-environment relationships in the
middle Caquetá area, Colombia, NW Amazonia. Vegetatio 120: 91-113.
———. 1996. Patterns of tree species richness in rainforests of the middle Caquetá area, Colombia, NW
Amazonia. Biotropica 28: 142-158.
———. & J. M. Lips. 1995. A land-ecological study of soils, vegetation, and plant diversity in
Colombian Amazonia. The Tropenbos Foundation, Wageningen, The Netherlands.
———, H. Balslev, J. Caveilier, C. Grandez, H. Tuomisto & R. Valencia. 2001. Evaluacion de
recursos vegetales no maderables en la Amazonia noroccidental. Institute for Biodivereity and
Ecosystem dynamics, Universiteit van Amsterdam, The Netherlands.
Dumont, J. F., S. Lamotte & F. Kahn. 1990. Wetland and upland forest ecosystems in Peruvian
Amazonia: plant species diversity in the light of some geological and botanical evidence. Forest
Ecology and Management 33/34: P125-139.
Encarnacion, F. 1993. El bosque y las formaciones vegetales en la llanura amazonica del Peru. Alma
Mater 6: 95–114.
Fine, P. V. A., Z. J. Miller, I. Mesones, S. Irazuzta, H. M. Apel, M. H. H. Stevens, I. Saaksjarvi, J. C.
Schultz & P. D. Coley. 2006. The growth-defense trade-off and habitat specialization by plants in
Amazonian forests. Ecology 87: S150-S162.
Ferreira, L. V. 2000. Effects of flooding duration on species richness floristic composition and forest
structure in river margin habitat in Amazonian blackwater floodplain forests: implications for future
design of protected areas. Biodiversity and Conservation 9: 1-14.
——— & G. T. Prance. 1998. Species richness and floristic composition in four hectares in teh Jau
National park in upland forests in central Amazonia. Biodiversity and conservation 7: 1349-1364.
——— & T. J. Stohlgren. 1999. Effects of river level fluctuation on plant species richness, diversity, and
distribution in a floodplain forest in Central Amazonia. Oecologia 120: 582–587.
———, M. T. F. Piedade & L. C. Bonates. 2006. Seed germination and seedling survival of
Himatanthus sucuuba (Spruce) Wood., in response to flooding in the varzeas of the Central Amazon.
Acta Amazonia 36: 413-418.
Fujisaka, S., G. Escobar & E. Veneklaas. 1998. Plant community diversity relative to human land uses
in an Amazon forest colony. Biodiversity and conservation 7: 41-57.
Gentry, A. H. 1986. Sumario de patrones fitogeograficos neotropicales y sus implicaciones para el
desarrollo de la Amazonia. Rev. Acad. Colomb. Cienc. Exactas. Fis. Nat. 16: 101-116.
———. 1988. Tree species richness of upper Amazonian forests. Proceedings of the National Academy of
Science. 85: 156-159.
Godoy, J. R., G. Petts & J. Salo. 1999. Riparian flooded forests of the Orinoco and Amazon basins: a
comparative review. Biodiversity and Conservation 8: 551-586.
Gorchov, D. L., F. Cornejo, C. Ascorra & M. Jaramillo. 1993. The role of seed dispersal in the natural
regeneration of rain forest after strip-cutting in the Peruvian Amazon. Vegetatio 107/108: 339-349.
———, J. M. Palmeirim, M. Jaramillo & C. F. Ascorra. 2004. Dispersal of seeds of Hymenaea
courbaril (Fabaceae) in a logged rain forest in the peruvian Amazonian. Acta Amazonia 34:
271-279.
Guillaument, J. L. 1987. Some structural and floristic aspects of the forest. Experientia 43: 241-251.
Haase, R. & S. G. Beck. 1989. Structure and composition of savanna vegetation in Northern Bolivia: a
preliminary report. Brittonia 41: 80-100.
288 R.W. Myster

Holdridge, L. R. 1967. Life zone Ecology. Tropical Science Center, San Jose Costa Rica.
Holling, C. S. 1992. Cross-scale morphology, geometry, and dynamics of ecosystems. Ecological
Monographs 62: 447-502.
Honorio, E. N. 2006. Floristic relationships of the tree flora of Jenaro Herrera, an unusual area of the
Peruvian Amazon. M.S. thesis, University of Edinburgh, Edinburgh, UK.
Junk, W. J. 1984. Ecology of the Varzea, floodplains of Amazonian white-water rivers. Pp. 215-243 in W.
T. Junk (ed.) The Amazon: limnology and landscape ecology of a mighty tropical river and its basin.
Kluwer, Dordrecht.
———. 1989. Flood tolerance and tree distribution in central Amazonian floodplains. Pp. 47-64 in L. B.
Holm-Nielsen, I. C. Nielsen, & H. Balslev (eds), Tropical forests: Botanical dynamics, speciation and
diversity. Academic, New York.
Kahn, F. & K. Mejia. 1990. Palm communities in wetland forest ecosystems of Peruvian Amazonia.
Forest Ecology and Management 33/34: 169-179.
Kalliola, R. S., M. Jukka, M. Puhakka & M. Rajasilta. 1991. New site formation and colonizing
vegetation in primary succession on the western Amazon floodplains. Journal of Ecology 79: 877-901.
Keller, M., A. Alencar, G. P. Asner, B. Braswell, M. Bustamante, E. Davidson, T. Feldpausch, E.
Fernandes, M. Goulden, P. Kabat, B. Kruijt, F. Luizao, S. Miller, D. Markewitz, A. D. Nobre, C.
A. Nobre, N. P. Filho, H. Da Rocha, P. S. Dias, C. Von Randow & G. L. Vourlitiis. 2004.
Ecological research in the large-scale biosphere atmosphere experiment in Amazonia: early results.
Ecological Applications 14: S3-S16.
Klinge, H., W. J. Junk & C. J. Revilla. 1990. Status and distribution of forested wetlands in tropical
South America. Forest Ecology and Management 33/34: 81-101.
Korning, J., K. Thomsen & B. Ollgaard. 1990. Composition and structure of a species rich Amazonian
rain forest obtained by two different sample methods. Nordic Journal of Botany 11: 103-110.
——— & H. Balslev. 1994. Growth and mortality of trees in Amazonian tropical rain forest in Ecuador.
Journal of Vegetation Science 4: 77-86.
Kubitzki, K. & A. Ziburski. 1994. Seed dispersal in floodplain forests on Amazonia. Biotropica 26: 30-
43.
Kvist, L. P. & G. Nebel. 2001. A review of Peruvian flood plain forests: ecosystems, inhabitants and
resource use. Forest Ecology and Mnaagement 150: 3-26
Lamotte, S. 1990. Fluvial dynamics and succession in the Lower Ucayali River basin, Peruvian
Amazonia. Forest Ecology and Management 33: 141-156.
Leimbeck, R. M. & H. Balslev. 2001. Species richness and abundance of epiphytic Araceae on adjacent
floodplain and upland forest in Amazonia Ecuador. Biodiversity and Conservation 10: 1579-1593.
Lieberman, M. & D. Lieberman. 1989. Forests are not just swiss cheese: canopy stereogeometry of non-
gaps in tropical forests. Ecology 70: 550-552.
Lopez, O. R. & T. A. Kursar. 1999. Flood tolerance of four tropical tree species. Tree Physiology 19:
925-932.
——— & T. A. Kursar. 2003. Does flood tolerance explain tree species distribution in tropical seasonally
flooded habitats? Oecologia 136: 193-204.
Losos, E. 1995. Habitat specificity of two palm species: experimental transplantation in Amazonian
successional forests. Ecology 76: 2595-2606.
Mayle, F. E., R. P. Langstroth, R. A. Fisher & P. Meir. 2007. Long-term forest- savannah dynamics in
the Bolivian Amazon: implications for conservation. Philosophical transactions of the Royal Society
B. 362: 291-307.
Myers, N., R. A. Mittermeier, C. G. Mittermeier, G. A. B. Da-Fonseca & J. Kent. 2000. Biodiversity
hotspots for conservation priorities. Nature 403: 853-858.
Myster, R. W. 2001. Mechanisms of plant response to gradients and after disturbances. The Botanical
Review 67: 441-452.
———. 2007a. Post-agricultural succession in the Neotropics. Springer-Verlag, NY.
———. 2007b. Interactive effects of flooding and forest gap formation on composition and abundance in
the Peruvian Amazon. Folia Geobotanica 42: 1-9.
Myster, R. W. & S. T. A. Pickett. 1988. Individualistic patterns of annuals and biennials in early
successional old fields. Vegetatio 78: 53-60.
Myster, R. W. & D. A. Schaefer. 2003. Species and microsite effects on litter decomposition in a Puerto
Rican landslide. Community Ecology 4: 157-162.
——— & P. G. Santacruz. 2005. “Una comparación de campo de insectos de suelo-morar de Amazonas:
Tierra firme y bosques de tierras inundadas vs. Espacios abiertos en el Parque Nacional Yasuní,
Ecuador.” Revista de la Pontificia Universidad Católica der Ecuador 76: 111-124.
Plant Communities of Western Amazonia 289

Nascimento, H. E. M. & W. F. Laurance. 2002. Total aboveground biomass in central Amazonian


rainforests: a landscape-scale study. Forest ecology and management 168: 311-321.
Nebel, G., J. Dragstead & J. K. Vanclay. 2001a. Structure and floristic composition of flood plain
forests in the Peruvian Amazon II. The understorey of restinga forests. Forest ecology and
Management 150: 59-77.
———, L. P. Kvist, J. K. Vanclay & H. Vidaurre. 2001b. Forest dynamics in floodplain forests in the
Peruvian Amazon: effects of disturbance and implications for management. Forest ecology and
Management 150: 79-92.
———, J. Dragsted & A. S. Vega. 2001c. Litterfall, biomass and net primary production in flood plain
forests in the Peruvian Amazon. Forest ecology and Management 150: 93-102.
———, L. P. Kvist, J. K. Vanclay, H. Christensen, L. Freitas & J. Ruiz. 2001d. Structure and floristic
composition of flood plain forests in the Peruvian Amazon I. Overstorey. Forest ecology and
Management 150: 27-57.
Neff, T., R. M. Lucas, J. R. Dos-Santos, E. S. Brondizio & C. C. Freitas. 2006. Area and age of
secondary forests in Brazilian Amazonia 1978-2002: An empirical estimate. Ecosystems 9: 609-623.
Normand, S., J. Vormisto, J. Svenning, C. Grandez & H. Balslev. 2006. Geographical and
environmental controls of palm beta diversity in paleo-riverine terrace forests in Amazonian Peru.
Plant Ecology 186: 161-176.
Notman, E., D. L. Gorchov & F. Cornejo. 1996. Effect of distance, aggregation, and habitat on levels of
seed predation for two mammal-dispersed neotropical rain forest tree species. Oecologia 106: 221-
227.
Padoch, C., J. M. Ayers, M. Pinedo-Vasquez & A. Herdrm. 1999. Varzea: Diversity, development, and
conservation of Amzzonia’s white water floodplains. NY Botanical Garden Press, Bronx, NY.
Parodi, J. L. & D. Freitas. 1990. Geographical aspects of forested wetlands in the lower Ucayali,
Peruvian Amazonia. Forest Ecology and Management 33/34: 157-168.
Parolin, P. 2000a. Growth, productivity and use of trees in white water floodplains. Pp. 375-391 in W. J.
Junk, J. J. Ohly, M. T. F. Piedade, & M. G. M. Soares (eds.) The Central Amazon floodplain: actual
use and options for a sustainable management. Backkhuys Publishers, Leiden.
———. 2000b. Seed mass in Amazonian floodplain forests with contrasting nutrient supplies. Journal of
Tropical Ecology 16: 417-428.
———. 2001a. Seed germination and early establishment of tree species from nutrient -rich and nutrient-
poor central Amazonian floodplains. Aquatic botany 70: 89-103.
———. 2001b. Morphological and physiological adjustments to waterlogging and drought in seedlinds of
Amazonian floodplain trees. Oecologia 128: 326-335.
———, W. J. Junk & M. T. F. Piedade. 2001. Gas exchange of six tree species from Central Amazonian
floodplains. Tropical Ecology 42: 15-24.
———. 2002a. Radial gradients in wood specific gravity in trees of central Amazonian floodplains. IAWA
Journal 23: 449-457.
———. 2002b. Life history and environment of Cecropia latiloba in Amazonian floodplains. Rev. Biol.
Trop. 50: 531-545.
———. 2002c. Seasonal changes of specific leaf mass and leaf size in trees of Amazonian floodplains.
Phyton 42: 169-185.
———. 2002d. Submergence tolerance vs. escape from submergence: two strategies of seedling
establishment in Amazonian floodplains. Environmental and experimental botany 48: 177-186.
———. 2003 Fugitive and possessive establishment strategies in Amazonian floodplain pioneers. Flora
198: 436-443.
———, A. C. Oliveira, M. T. F. Piedade, F. Wittmann & W. J. Junk. 2002a. Pioneer trees in
Amazonian floodplains: three key species form Monospecific stands in different habitats. Folia
Geobotanica 37: 225-238.
———, N. Armbruster & W. J. Junk. 2002b. Seasonal changes of leaf nitrogen content in trees of
Amazonian floodplains. Acta Amazonia 32: 231-240.
———, L. V. Ferreira & W. J. Junk. 2003. Germination characteristics and establishment of trees from
central Amazonian flood plains. Tropical Ecology 44: 155-167.
———, J. Adis, M. F. Da-Silva, L. I. Do-Amaral, L. Schmidt & M. T. F. Piedade. 2003. Floristic
composition of a floodplain forest in teh Anavilhanas archipelago, Brazilian Amazonia. Amazoniana
17: 399-411.
———, O. DeSimone, K. Haase, D. Waldhoff, S. Rottenberger, U. Kuhn, J. Kesselmeier, B. Kleiss,
W. Schmidt, M. T. F. Piedade & W. J. Junk. 2004. Central Amazonian floodplain forests: tree
adaptations in a pulsing system. The Botanical Review 70: 357-380.
290 R.W. Myster

Peres, C. A. 1994. Composition, density and fruiting phenology of arborescent palms in an Amazonian
terra firme forest. Biotropica 26: 285-294.
Phillips, O. L., R. V. Martinez, L. Arroyo, T. R. Baker, T. Killeen, S. L. Lewis, Y. Malhi, A. M.
Mendoza, D. Neill, P. N. Vargas, M. Alexiades, C. Ceron, A. Di Fiore, T. Erwin, A. Jardim, W.
Palacios, M. Saldias, & B. Vincetti. 2002. Increasing dominance of large lianas in Amazonian
forests. Nature 418: 770-774.
Pires, J. M. & G. T. Prance. 1985. The vegetation types of the Brazilian Amazon. Pp 109-145. In: G. T.
Prance, and T. E. Lovejoy, eds. Amazonia. Pergamon press, Oxford.
Pitman, N. C. A., J. Terborgh, M. R. Silman, & P. Nunez. 1999. Tree species distributions in an upper
Amazonian forest. Ecology 80: 2651-2661.
———, J. W. Terborgh, M. R. Silman, P. V. Nunez, D. A. Neill, C. E. Ceron, W. A. Palacios, & M.
Aulestia. 2001. Dominance and distribution of tree species in upper Amazonian terra firme forests.
Ecology, 82: 2101-2117.
Poorter, L. & Y. Hayashida-Oliver. 2000. Effects of seasonal drought on gap and understorey seedlings
in a Bolivian moist forest. Journal of Tropical Ecology 16: 481-498.
Poulsen, A. D. & H. Balslev. 1991. Abundance and cover of ground herbs in an Amazonian rain forest.
Journal of Vegetation Science 2: 315-322.
———, H. Tuomisto, & H. Balslev. 2006. Edaphic and floristic variation within 1-ha plot of lowland
Amazonian rain forest. Biotroica 38: 468-478.
Prance, G. T. 1979. Notes on the vegetation of Amazonia III. The terminology of Amazonian forest types
subject to inundation. Brittonia 31: 26-38.
Puhakka, M., R. Kalliola, M. Rajasilta, & J. Salo. 1992. River types, site evolution and successional
vegeation patterns in Peruvian Amazonia. Journal of Biogeography 19: 651-665.
Rice, A. H., F. H. Pyle, S. R. Saleska, L. Hutyra, M. Palace, M. Keller, P. B. De-Camargo, K.
Portilho, D. F. Marques, & S. C. Wofsy. 2004. Carbon balance and vegetation dynamics in an old-
growth Amazonian forest. Ecological Applications 14: S55-S77.
Romo, M., H. Tuomisto, & B. A. Loiselle. 2004. On the density-dependence of seed predation in
Dipteryz micrantha, a bat-dispersed rain forest tree. Oecologia 140: 76-85.
Ruokolainen, K., H. Tuomisto, M. J. Macia, M. A. Higgins, and M. Yli-Halla. 2007. Are floristic and
edaphic patterns in Amazonian rain forests congruent for trees, pteridophytes and Melastomataceae?
Journal of Tropical Ecology 23: 13-25.
Saldarriaga, J. G., D. C. West, M. L. Tharp, & C. Uhl. 1988. Long-term chronosequence of forest
succession in the upper rio negro of Colombia and Venezuela. Journal of Ecology 76: 938-958.
Salo, J., R. Kalliola, I. Hakkinen, Y. Makinea, P. Niemela, M. Puhakka, & P. D. Coley. 1986. River
dynamics and the diversity of Amazon lowland forest. Nature 322: 254-258.
Sanford, R. L. 1989. Root systems of three adjacent old growth Amazon forests and associated transition
zones. Journal of Tropical Forest Science 3: 268-279.
Shukla, J., C. Nobre, & P. Sellers. 1990. Amazon deforestation and climate change. Science 247
(4948):1322-1325.
Soares-Filho, B. S., D. C. Nepstad, L. M. Curran, G. C. Cerqueira, R. A. Garcia, C. A. Ramos, E.
Voll, A. McDonal, P. Lefebvre, & P. Schlesinger. 2006. Modelling conservation in the Amazon
basin. Nature 440/23: 520-523.
Spichiger, R., P. Loizeau, C. Latour & G. Barriera. 1996. Tree species richness of a South-Western
Amazonian forest (Jenaro Herrera, Peru, 73o40’W/4o54’S). Candollea 51: 559-577.
Svenning, J. 1999. Microhabitat specialization in a species-rich palm community in Amazonian Ecuador.
Journal of Ecology 87: 55-65.
Terborgh, J. & E. Andresen. 1998. The composition of Amazonian forests: patterns at local and regional
scales. Journal of Tropical Ecology 14: 645-664.
Tuomisto, H., K. Ruokolainen, A. D. Poulsen, R. C. Moran, C. Quintana,G. Canas, & J. Cell. 2002.
Distribution and diversity of Pteridophytes and Melastomataceae along Edaphic gradients in Yasuni
National Park, Ecuadorian Amazonia. Biotropica 34: 516-533.
———, K. Ruokolainen, & M. Yli-Halla. 2003. Dispersal, environment, and floristic variation of
Western Amazonian forests. Science 299: 241-244.
Uhl, C. & P. G. Murphy. 1981. Composition, structure and regeneration of a Terra firme forest in the
Amazon basin of Venezuela. 1981. Tropical ecology 22: 219-237.
Valencia, R., H. Balslev, G. Paz & C. Mino. 1994. High tree alpha-diversity in Amazonian Ecuador.
Biodiversity and Conservation 3: 21-28.
Plant Communities of Western Amazonia 291

———, R. B. Foster, G. Villa, R. Condit, J. Svenning, C. Hernandez, K. Romoleroux, E. Losos,


E. Magard, & H. Balslev. 2004. Tree species distributions and local habitat variation in the Amazon:
large forest plot in eastern Ecuador. Journal of Ecology 92: 214-229.
Vormisto, J., J. Svenning, P. Hall, & H. Balslev. 2004. Diversiy and dominance in palm (Arecaeae)
communities in terra firme forests in the western Amazon basin. J. Ecology 92: 577-588.
Walter, H. 1973. Vegetation of the earth and the ecological systems of the geo -biosphere. Springer-
Verlag. New York.
Watt, A. S. 1947. Pattern and processes in the plant community. Journal of Ecology 35: 1-22.
Whittaker, R. H. 1975. Communities and ecosystems. MacMillian, New York.
Wilson, J. B. 1991. Does vegetation science exist? Journal of Vegetation Science 2: 289-290.
Wittmann, F. & W. J. Junk. 2002. Sapling communities in Amazonian white-water forests. Journal of
Biogeography 30: 1533-1544.
———, D. Anhuf, & W. J. Junk. 2002. Tree species distribution and community structure of central
Amazonian varzea forests by remote-sensing techniques. Journal of Tropical Ecology 18: 805-820.
———, W. J. Junk, & M. T. F. Piedade. 2004. The varzea forests in Amazonia: flooding and the highly
dynamic geomorphology interact with natural forest succession. Forest ecology and Management 196:
199-212.
Worbes, M. 1997. The forest ecosystem of the floodplains. Pp 223-265 in W. J. Junk (ed) The central
Amazon floodplains: ecology of a pulsing system, Ecological studies 126. Springer, New York.
———, H. Klinge, J. D. Revilla, & C. Martius. 1992. On the dynamics, floristic subdividion and
geographical distribution of varzea forests in Central Amazonia. Journal of Vegetation Science 3: 553-
564.
Wyatt, J. L. & M. R. Silman. 2004. Distance-dependence in two Amazonian palms: effects of spatial and
temporal variation in seed predator communities. Oecologia 140: 26-35.

You might also like